You are on page 1of 13

JOM

https://doi.org/10.1007/s11837-021-05032-y
 2021 The Minerals, Metals & Materials Society

SOLID FREEFORM FABRICATION 2021

Microstructure and Deformation Behavior of Additively


Manufactured 17–4 Stainless Steel: Laser Powder Bed Fusion
vs. Laser Powder Directed Energy Deposition

P.D. NEZHADFAR,1,2 PAUL R. GRADL,3 SHUAI SHAO,1,2


and NIMA SHAMSAEI 1,2,4

1.—Department of Mechanical Engineering, Auburn University, Auburn, AL 36849, USA.


2.—National Center for Additive Manufacturing Excellence (NCAME), Auburn
University, Auburn, AL 36849, USA. 3.—Propulsion Department, NASA Marshall Space Flight
Center, Huntsville, AL 35812, USA. 4.—e-mail: shamsaei@auburn.edu

This study aims to compare the microstructure of 17–4 PH stainless steel (SS)
manufactured via laser powder bed fusion (L-PBF) and laser powder directed
energy deposition (LP-DED) in non-heat treated (NHT) and heat treated
conditions. In addition, the room-temperature tensile behavior of heat-treated
L-PBF and LP-DED 17–4 PH SS has been investigated and compared with
that of the wrought counterpart with the same heat treatment conditions. The
results show that the L-PBF specimens have a finer microstructure (ferrite +
lath martensite) than the LP-DED ones (massive ferrite + Widmanstätten
ferrite) in NHT condition. Electron backscatter diffraction analysis shows that
the L-PBF and LP-DED specimens have twin-based substructure lath
martensite after heat treatment. Despite the lower tensile strength of the LP-
DED specimens compared with the L-PBF ones, the ductility of peak-aged LP-
DED specimens was reduced due to the presence of the d-ferrite phase having
a significant plastic deformation incompatibility with the martensite.

Among the various AM processes, laser powder


INTRODUCTION
bed fusion (L-PBF) and laser powder directed
Additive manufacturing (AM) has transformed energy deposition (LP-DED) are the most prominent
the manufacturing process of structural parts in laser-based processes, which have been extensively
various industries such as aerospace, automotive, investigated in the literature.3 The L-PBF process
biomedical, defense, and nuclear. The layer-by-layer provides more freedom in designing near-net-
and track-by-track nature of AM processes allows shaped parts with higher geometrical resolution
the manufacture of near-net-shaped parts with and precision than the LP-DED process; however,
complex geometries, reduces the cost and lead L-PBF is typically limited to a single-powder feed-
times, and facilitates the fabrication of highly stock.4 On the other hand, LP-DED offers better
customized parts for specific applications (e.g., in compatibility for multi-powder feedstock, function-
the medical field).1 There are various types of AM ally graded components, and the fabrication of large
techniques for metallic materials classified based on parts.1
feedstock form (e.g., powder versus wire), feeding One of the specific characteristics of the laser-
mechanism (e.g., powder bed versus blown powder), based AM processes is the high heating/cooling
and energy source (e.g., laser versus electron rates during fabrication, resulting in a refined
beam).2 microstructure compared to conventionally manu-
factured (CM) counterparts. Accordingly, the AM
materials often have static mechanical properties
comparable to, and in some cases outperforming,
their CM counterparts.5,6 The mechanical perfor-
mance of the AM materials is primarily influenced
(Received September 30, 2021; accepted November 5, 2021)
Nezhadfar, Gradl, Shao, and Shamsaei

by the material’s structure (i.e., grain structure, finer grain structure, reduced the agglomeration of
texture, surface roughness, and defect structure), the strengthening particles (i.e., Y2O3), and yielded
which is impacted by the thermal history induced higher hardness values in the ODS steel.
during the manufacturing. The thermal history The significant dependence of the microstructure
itself is governed by the process parameters (i.e., and mechanical properties of such known materials
powder, scan speed, etc.) and design parameters as IN718, Ti-6Al-4V, ODS steels, etc. on AM pro-
(i.e., part size and geometry, time interval, etc.). In cesses necessitates the careful evaluation of this
addition to the above-mentioned influential param- dependence before mass adoption of AM to any
eters, the AM process technique (e.g., L-PBF, LP- other materials in key industrial sectors. The 17-4
DED, etc.) itself may also cause variations in the precipitation hardening (PH) stainless steel (SS) is
microstructural and mechanical properties of a one of the most commonly used materials in various
material system. industries (e.g., aerospace, energy, food processing,
The LP-DED process, due to the slower moving etc.) due to its favorable chemical and mechanical
melt pool (typically 10–30 mm/s7 compared to properties (e.g., high corrosion resistance, high
1000–1700 mm/s for L-PBF8) created by a high strength, and ductility, etc.).18 It has recently
powered laser (typically 1000 W), has a consider- drawn much attention from the AM community
ably lower cooling rate (by approximately three due to its superior weldability.19,20 The sensitivity of
orders of magnitude) than L-PBF,9,10 giving the its microstructure to the cooling rate during solid-
fabricated materials a distinct microstructure (i.e., ification, as well as the subsequent thermal expo-
grain structure, crystallographic texture, precipita- sures, can potentially make the properties of 17–4
tions, etc.) and defect content. Amato et al.11 PH SS process-dependent. This study investigates
reported the presence of a spheroidal/ellipsoidal c‘‘ and compares the microstructure and room-temper-
in c matrix of non-heat treated (NHT) L-PBF IN718, ature tensile behavior of 17–4 PH SS manufactured
while Laves phase has been characterized in an by the L-PBF and LP-DED processes. The grain
interdendritic c matrix of NHT LP-DED counter- structure and phase constituent of the L-PBF and
parts12,13 Xu et al.14 reported a lamellar a+ß struc- LP-DED 17–4 PH SS specimens are characterized
ture in L-PBF Ti-6Al-4V, whereas Caroll et al.15 and compared in the NHT condition. In addition,
showed a Widmanstätten structure along with the effect of various heat treatment conditions on
grain boundary a phase in the coarse columnar the crystallographic texture and room-temperature
prior b grains in LP-DED Ti-6Al-4V. It has been tensile behavior of the L-PBF and LP-DED 17–4 PH
reported that even post-thermal heat treatment SS specimens is investigated.
may not completely diminish the differences in the
microstructure of the L-PBF and LP-DED AM EXPERIMENTAL
materials. Schneider et al.16 reported different
Two batches of argon-atomized 17–4 PH SS
grain structures (i.e., size and morphology) for L-
powder were used to fabricate the specimens via
PBF and LP-DED IN718 even after applying a
the L-PBF and LP-DED processes, respectively. The
similar heat treatment procedure, which was attrib-
detailed chemical compositions of the 17–4 PH SS
uted to the variation in their initial microstructure
powders used for each AM process are listed in
(i.e., NHT condition). The L-PBF specimens showed
Table I.
the highest degree of homogenization and more
An EOS M290 machine was used to fabricate the
refined grains than the LP-DED ones.
L-PBF cylindrical bars (/11 mm). The EOS default
The microstructure variations of the L-PBF and
process parameters, laser power of 220 W, scan
LP-DED AM materials result in different mechan-
speed of 755 mm/s, hatching distance of 0.1 mm, and
ical responses. Babuska et al.4 recently reported a
layer thickness of 0.04 mm, were employed, and
higher tensile strength for L-PBF CoCr specimens
nitrogen was used as the shielding gas. The LP-
than for LP-DED counterparts. They attributed this
DED cylindrical bars (/15.24 mm) were fabricated
to the finer microstructure of the L-PBF specimens
using an RPM Innovations (RPMI) 557 machine.
due to the higher cooling rate compared with the
The process parameters were: power of 1070 W,
LP-DED ones. Donate-Buendia et al.17 compared
layer height of 0.38 mm, travel speed of 1,016 mm/
the microstructure, and mechanical properties of
min, and powder feed rate of 15.1 g/min, with argon
oxide dispersion strengthened (ODS) steels manu-
used as the shielding gas.
factured via L-PBF and LP-DED processes. Com-
pared to LP-DED, the L-PBF process produced a

Table I. Chemical composition of 17-4 PH SS powders used for fabrication

C Cr Ni Cu Mn Si Nb Mo N O P S Nb+Ta Fe

L-PBF (Wt.%) 0.01 15.80 4.60 3.67 0.51 0.32 0.32 0.21 0.02 0.04 0.035 0.014 0.27 Bal.
LP-DED (Wt.%) 0.01 16.39 4.17 3.32 0.06 0.78 0.27 0.04 0.01 0.02 0.00 0.00 0.27 Bal.
Microstructure and Deformation Behavior of Additively Manufactured 17–4 Stainless Steel:
Laser Powder Bed Fusion vs. Laser Powder Directed Energy Deposition

The Thermo-Calc. software was used to generate DED 17-4 PH SS. All the L-PBF and LP-DED
the Fe-Cr binary phase diagram to analyze the specimens underwent hot isostatic pressing (HIP) at
phase transformations possible for 17-4 PH SS. The 1163C/3 h under 103 MPa pressure and were
TCFE9 thermodynamic database for various Fe- solution (Sol) treated at 1050C/0.5 h [i.e., Condition
based alloys and steels such as stainless steels was A (CA)] followed by air cooling. Some specimens for
employed.21,22 In addition, to incorporate the influ- each type (i.e., L-PBF and LP-DED specimens) were
ence of all the alloy elements in generating the aged at 482C/1 h (i.e., H900) known as peak-age,
phase diagram, the Nieq (nickel equivalent) and Creq and some were over-aged at 621C/4 h (i.e., H1150).
(chromium equivalent) were obtained following the In the following text, the heat-treated specimens
Schaeffler equations:23 (i.e., SR + HIP + Sol + Age) are dubbed CA-H900
and CA-H1150. After the full heat treatment, all the
Nieq ðwt%Þ ¼ %Ni þ 0:5ð%MnÞ þ 0:3ð%CuÞ specimens were machined to the final tensile testing
þ 25ð%NÞ þ 30ð%CÞ ð1Þ geometry following the ASTM E8 standard.25
The L-PBF and LP-DED specimens were cut in
the gage section transversely parallel to the build
Creq ðwt%Þ ¼ %Cr þ 2ð%SiÞ þ 1:5ð%MoÞ direction for microstructure characterization prior
þ 1:75ð%NbÞ ð2Þ to tensile testing. The microstructure was charac-
terized in NHT and fully heat treated conditions.
The Nieq and Creq values were further calculated The samples were ground and polished using sand-
for the L-PBF to be 7.6 wt% and 16.6 wt%, respec- papers with grits 320–4000 followed by a mirror-
tively. For the LP-DED 17-4 PH SS, the Nieq and finish polishing step using Chemo-met along with a
Creq values were 6.1 and 17.7 wt%, respectively. 0.2-lm colloidal silica suspension. The microstruc-
Figure 1 shows the binary phase diagram generated ture was further characterized on the plane parallel
by Thermo-Calc. software; the Cr amounts for the L- (XZ-plane) to the build direction using a Zeiss
PBF and LP-DED specimens are indicated on the Crossbeam 550 scanning electron microscope
phase diagram to predict the phase constituents at (SEM) with an electron backscatter diffraction
various temperatures. According to the phase dia- (EBSD) detector. The specimens were mirror-fin-
gram, the stress relieve (SR) heat treatment was ished using a vibratory polisher for 2 h before
carried out at 650C/1 h for the L-PBF and LP-DED conducting the EBSD analysis and electron chan-
specimens. It has been reported in Ref. 24 that this neling contrast imaging (ECCI).
SR procedure changes neither the microstructure
nor the mechanical properties of the L-PBF and LP-

Fig. 1. Ni-Cr binary phase diagram samples generated by Thermo-Calc. software using the TCFE9 thermodynamic database22.
Nezhadfar, Gradl, Shao, and Shamsaei

The tensile tests were conducted on the heat- Since the martensite start temperature (i.e., Ms)
treated L-PBF and LP-DED 17-4 PH SS specimens is above room temperature for the 17–4 PH SS
at room temperature at 0.005 mm/mm/min strain (Ms = 100–150C),26,27 the austenite formed during
rate. The fracture surfaces were further cleaned in a solidification transforms to the martensite upon
bath of isopropanol and water using an ultrasonic cooling to room temperature. However, the ferritic
cleaner before performing fractography using SEM. microstructure of the NHT AM 17-4 PH SS speci-
mens, regardless of the manufacturing technique,
RESULTS AND DISCUSSION has been attributed to the ‘‘austenite by-passing’’
mechanism. In this mechanism, the ferrite phase
NHT Microstructure
formed from the liquid will not transfer to the
The NHT microstructures of the L-PBF and LP- austenite due to the high cooling rate in AM
DED 17-4 PH SS specimens characterized via EBSD processes as compared to conventional manufactur-
analysis and ECCI are shown in Fig. 2. It is evident ing techniques (e.g., casting, forging, etc.).28
that the L-PBF 17-4 PH SS specimen (Fig. 2a) has a The lath martensite in the microstructure of L-
more refined grain structure as compared to the LP- PBF 17-4 PH SS specimen (see Fig. 2a), however,
DED 17–4 PH SS one (Fig. 2b); the L-PBF specimen indicates that the austenite has been formed during
consists of fine, primarily equiaxed, ferrite grains, the solidification. This may be attributed to the N2
whereas the LP-DED one consists of very coarse shielding gas used to fabricate the L-PBF speci-
columnar ferrite grains. The ECCI images with mens. It has been reported that N2 as the shielding
higher magnifications reveal that the microstruc- gas refines the grain structure and also stimulates
ture of the L-PBF specimen consists of ferrite and austenite formation during solidification.20,29,30
lath martensite. In contrast, the LP-DED specimen Therefore, the refined ferrite grains (average grain
has a ferritic microstructure consisting of massive size of 6 lm) and the lath martensite in the
ferrite grains with Widmanstätten ferrites decorat- microstructure are the result of using N2 as shield-
ing the grain boundaries. ing gas. The coarse microstructure of the LP-DED

Fig. 2. Inverse pole figure (IPF) maps along the Z direction and ECCI micrographs of (a) L-PBF, and (b) LP-DED 17-4 PH SS specimens in NHT
condition.
Microstructure and Deformation Behavior of Additively Manufactured 17–4 Stainless Steel:
Laser Powder Bed Fusion vs. Laser Powder Directed Energy Deposition

17-4 PH SS specimen (average grain size of 93.1 Heat Treated Microstructure


lm) may be attributed to the lower cooling rate
The microstructure of the L-PBF and LP-DED 17-
induced in the LP-DED process as compared to the
4 PH SS specimens are compared in Fig. 3 for the
L-PBF one. Although the cooling rate of the LP-
CA-H900 and CA-H1150 heat treatment conditions.
DED process is believed to be above the threshold
Although there is a visible difference between the
for austenite by-passing, it has been reported to
microstructures of the NHT L-PBF and LP-DED 17-
have almost a two to three orders of magnitude
4 PH SS specimens (see Fig. 2), post-thermal treat-
slower cooling rate than the L-PBF process.4 In
ment (i.e., SR + HIP + HT) have evidently altered
addition, the ferrite phase is more stabilized in LP-
the microstructures to martensite-dominated ones
DED specimens due to the higher Creq/Nieq value for
for both the L-PBF and LP-DED specimens (see
the LP-DED 17-4 PH SS (2.9) compared to that of
Fig. 3). It can be seen in Fig. 3a and c that both the
the L-PBF counterpart (2.1). The significantly
L-PBF and LP-DED have a primarily martensitic
different NHT microstructures of the L-PBF and
microstructure in the CA-H900 heat treatment
LP-DED 17-4 PH SS specimens may not only lead to
condition with a minimal fraction of retained
drastically different mechanical properties but also
austenite (0.1%). The fraction of retained austen-
may be inferior in strength due to the absence of
ite is increased for both the L-PBF (Fig. 3b) and LP-
precipitates. In hope to resolve these issues, post-
DED (Fig. 3d) specimens that have undergone CA-
thermal treatments are typically performed to
H1150 heat treatment. This is due to the diffusion of
enhance the mechanical properties of the material,
austenite stabilizer elements (i.e., Cu, Ni, N, etc.) to
as well as reducing the differences in the
the grain boundaries during the long-term aging
microstructures of the L-PBF and LP-DED
(i.e., 4 h) at a temperature close to the austenite
specimens.
reversion (see Fig. 1), which may have resulted in
the nucleation of austenite grains.31 However,
depending on the size of the reverted austenite
grains, they may or may not transform to

Fig. 3. IPF maps (along the Z direction) and phase maps for the heat treated 17-4 PH SS specimens: (a) L-PBF and (c) LP-DED at CA-H900,
and (b) L-PBF and (d) LP-DED at CA-H1150. Note that the black boundaries in the magnified IPF maps (in the middle) represent the prior
austenite grain boundaries.
Nezhadfar, Gradl, Shao, and Shamsaei

martensite upon cooling. It has been well estab- compared for the L-PBF and LP-DED specimens in
lished that martensitic transformations are more both CA-H900 and CA-H1150 conditions.
difficult for finer austenite grains, which may be In addition to the ODF maps (0 £ U1, U, u2 £ 90),
retained after heat treatment.32 the fraction of texture components for the L-PBF
As seen in Fig. 3b for the CA-H1150-treated L- and LP-DED specimens which have undergone CA-
PBF specimen, the fraction of austenite (4.3%) H900 and CA-H1150 heat treatments are also
retained in the microstructure is higher than that of provided in Figs. 4c and 5c, respectively. In the
the CA-H1150 LP-DED counterpart (0.5%), shown CA-H900 heat treatment condition, the L-PBF
in Fig. 3d. This can be partially ascribed to the finer specimen (see Fig. 4a) has a lower texture intensity
initial microstructure of the L-PBF specimen com- [i.e., multiple random density (mrd)], mrdmax =
pared with that of the LP-DED one, which results in 1.85, than the LP-DED one, mrdmax = 3.25 (see
finer reverted austenite grains that are more diffi- Fig. 4b). This is likely due to the finer grain
cult to transform to martensite upon cooling. More- structure of the L-PBF specimen and most likely
over, using N2 shielding gas for fabrication may with more randomly oriented grains than LP-DED
increase the austenite stabilization and hinder the one. The texture intensities for both the L-PBF
martensitic transformation. It can be observed that and LP-DED specimens are reduced as compared
the d-ferrite phase is retained in the microstructure to their NHT conditions ((mrdmax)L-PBF = 2.03,
of the LP-DED specimens (see Fig. 3c and d), (mrdmax)LP-DED = 20.87) reported in24 and become
regardless of the heat treatment condition, whereas more randomized. This is ascribed to the recrystal-
the d-ferrite phase is absent in the microstructure of lization occurring during HIP and to further heat
the N2-shielded L-PBF specimens. The N2 shielding treatment procedures (i.e., Sol + Age). Therefore,
gas has been reported to reduce the d-ferrite phase full heat treatment could to some extent weaken
fraction retained in the microstructure of laser- texture differences between the L-PBF and LP-DED
welded materials at room temperature. N as an 17–4 PH SS specimens induced due to the differ-
austenite stabilizer results in the transformation of ences in the cooling rates of the AM processes.
the d-ferrite phase to austenite during the solidifi- The main texture component for the L-PBF
cation, which reduces the fraction of d-ferrite phase specimen at the CA-H900 heat treatment condition
retained in the microstructure at room temperature. is TB, whereas TC1 is the main texture component
Moreover, the L-PBF specimen has a lower Creq/ for the LP-DED specimen (see Fig. 4c). It can be
Nieq value than the LP-DED one, reducing ferrite seen in Fig. 5c that the conducting CA-H1150 heat
stability during solidification. treatment will result in a more similar trend of
texture component fractions in the L-PBF and LP-
Texture Analysis DED specimens as compared to the CA-H900 heat-
treated ones. This may be due to the longer duration
The variation in thermal history induced via L-
of aging close to the austenite reversion tempera-
PBF and LP-DED processes influences the crystal-
ture, and likely have resulted in partial reversion of
lographic texture of the AM 17–4 PH SS specimens.
the austenite. The reverted austenite then trans-
It has been shown that the NHT L-PBF 17–4 PH SS
forms to the martensite. The transformed-type
specimen possesses a strong cube along with weak c-
texture components represent the texture formed
fiber texture components, whereas no specific tex-
from the parent austenite phase (face-centered
ture component has been reported for the NHT LP-
cubic) during martensitic transformation.
DED 17-4 PH SS specimen due to its very large
Ping et al.33 have shown the formation mecha-
grain structure (see Fig. 2b).24 In addition, it has
nism of lath martensite with a twin substructure
been reported that performing SR heat treatment
with a {112}<111> relationship. Figure 6 shows the
assists with the recrystallization upon further heat-
{112} pole figures as well as the linear misorienta-
treatment procedures and weakens the texture.24
tion profiles along the arrows elongated through the
To understand and compare the crystallographic
selected prior austenite grain (PAG) for the L-PBF
textures of the L-PBF and LP-DED 17–4 PH SS
and LP-DED specimens which have undergone CA-
specimens after applying full heat treatments (i.e.,
H900 and CA-H1150. The poles coincidence in the
SR + HIP + Sol + Age), their corresponding orien-
{112} maps for all the conditions are in line with the
tation distribution function (ODF) maps for each
reported pole figures for martensite with the Kur-
condition are presented in Figs. 4 and 5. The ODF
djumov–Sachs (K-S) relationship with the parent
maps have been analyzed in parallel to the build
austenite.34 In addition, the point-to-point misori-
and loading direction (i.e., Z) using a spherical
entation profiles show the boundaries with 60
harmonics method, keeping the Euler angle 3 (i.e.,
misorientation indicating the twin boundaries,35
u2) constant. The quantified fraction of the essential
which further proves the twin-based martensitic
texture components, i.e., Cube (C), Rotated Cube
transformation in the L-PBF and LP-DED 17-4 PH
(RC), Copper (Cu), Transformed Copper (TC), Goss
SS specimens.
(G), Transformed Goss (TG), Brass (B), Trans-
formed Brass (TB), and c-fiber) are presented and
Microstructure and Deformation Behavior of Additively Manufactured 17–4 Stainless Steel:
Laser Powder Bed Fusion vs. Laser Powder Directed Energy Deposition

Fig. 4. ODF maps for the (a) L-PBF and (b) LP-DED 17-4 PH SS specimens, and (c) their quantified texture components for the CA-H900 heat
treatment condition. A schematic of the important texture components in body-centered cubic for u2 = 45 is also shown.

Tensile Behavior and Fractography Analysis (i.e., CA-H1150) significantly reduces the strength
and increases the ductility. The high tensile
The engineering stress-engineering strain curves
strength (i.e., Sy, Su) after the CA-H900 condition
for the L-PBF and LP-DED specimens which have
is attributed to the formation of nano-sized Cu-
undergone CA-H900 and CA-H1150 heat treatment
enriched precipitates in this heat treatment condi-
conditions are presented in Fig. 7a. The tensile
tion, while these precipitates are coarsened upon
properties (i.e., Sy, Su, and %El) for the L-PBF, LP-
over-aging resulting in higher ductility.36
DED, and wrought 17-4 PH SS are compared in
Comparing the L-PBF and LP-DED specimens,
Fig. 7b. In general, the peak-age heat treatment
the L-PBF specimens have higher tensile strength
(i.e., CA-H900) results in the highest tensile
than those of the LP-DED counterparts regardless
strength in 17–4 PH SS at the expense of the
of the heat treatment condition. This is because of
ductility, whereas the over-aging heat treatment
the finer microstructure obtained for the L-PBF
Nezhadfar, Gradl, Shao, and Shamsaei

Fig. 5. ODF maps for the (a) L-PBF and (b) LP-DED 17-4 PH SS specimens, and (c) the quantified texture components for the CA-H1150 heat
treatment condition. A schematic illustration of the important texture components in body-centered cubic for u2 = 45 is also shown.

specimens compared with that of the LP-DED ones of the expected ductility in the CA-H900 heat-
(see Fig. 3). The L-PBF specimen has higher tensile treated LP-DED specimen is due to the presence of
strength, whereas the LP-DED specimen possesses d-ferrite in its microstructure (see Fig. 3c). Figure 8
comparable tensile strength and ductility compared shows the kernel average misorientation (KAM),
with the wrought 17-4 PH SS. The higher tensile and the Schmid factor map of the body-centered
properties of the L-PBF 17-4 PH SS are ascribed to cubic slip system (i.e., {110} <111>) along the
its finer grain structure. loading direction (i.e., parallel to the Z axis) for
Although the CA-H900 heat treated LP-DED the LP-DED 17-4 PH SS specimens heat treated at
specimen has a lower tensile strength than the L- CA-H900 (Fig. 8a and b). The d-ferrite phase is
PBF counterpart and is expected to have higher indicated by arrows in the KAM images and circled
ductility, its ductility is comparable to that of the L- in the Schmid factor maps. As shown in Fig. 8a and
PBF specimen with higher tensile strength. The loss b for the CA-H900 heat-treatment condition, the d-
Microstructure and Deformation Behavior of Additively Manufactured 17–4 Stainless Steel:
Laser Powder Bed Fusion vs. Laser Powder Directed Energy Deposition

Fig. 6. IPF maps, {112} pole figures, and the point-to-point misorientation maps showing the twin-based substructure of the martensite in
selected PAG of the L-PBF and LP-DED 17-4 PH SS specimens. CA-H900 condition: (a) L-PBF, (b) LP-DED, and CA-H1150 condition: (c) L-
PBF, and (d) LP-DED.

ferrite phase not only has well-defined edges but it The fracture surfaces of the L-PBF and LP-DED
also has a strong contrast in the Schmid factor with 17-4 PH SS specimens are compared in Figure 9a
the martensitic matrix; in other words, there is a for CA-H900 and Fig. 9b for the CA-H1150 heat-
high incompatibility between the d-ferrite and the treatment conditions. Both conditions result in
matrix resulting in loss of ductility considering that mostly axis-symmetric fracture surfaces with the
the material possesses high strength. shear lips of CA-H900 condition occupying very
small area fractions and those of CA-H1150
Nezhadfar, Gradl, Shao, and Shamsaei

Fig. 7. Tensile behaviors of the L-PBF and LP-DED 17-4 PH SS specimens which have undergone CA-H900 and CA-H1150 heat treatment
conditions: (a) flow stress curves, (b) summarized tensile properties, and (c) %elongation versus ultimate tensile strength. The wrought data has
been. taken from ASTM A69337

Fig. 8. Strain analysis of the d-ferrite phase for the LP-DED 17-4 PH SS specimen which has undergone CA-H900 heat treatment condition: (a)
KAM, (b) Schmid factor maps. Note the loading direction parallel to the Z-axis considered for generating the Schmid factor map.

condition being large. The very well-defined cup- shearing.38 The CA-H900-treated LP-DED speci-
and-cone feature CA-H1150 fracture surfaces men has larger cracks and facets than the L-PBF
(Fig. 9b) suggests that their fracture was very ones, which resulted in the loss of ductility despite
ductile. This is consistent with the presence of of having lower strength. This may be related to the
quasi-cleavage facets and cracks on the fracture presence of large d-ferrite in the microstructure of
surfaces of the CA-H900 heat treated specimens the LP-DED specimen, having high incompatibility
showing brittle behavior, while enormous dimples with the matrix, which resulted in crack initiation
formed on the fracture surfaces of the CA-H1150 and growth along the lath martensite boundaries.
heat treated specimens representing ductile On the other hand, CA-H1150 treated LP-DED
behavior. specimens have larger and deeper dimples due to
Another main difference between the fracture the coarser microstructure, justifying its higher
surfaces of the L-PBF and LP-DED specimens is the ductility compared with the L-PBF ones.
presence of cracks in the CA-H900 condition, and
dimples in the CA-H1150 condition. This suggests CONCLUSION
that the fracture surfaces of the former condition
This study characterized and compared the non-
had radial zones which were evidence of rapid crack
heat treated (NHT) microstructures of 17–4 PH SS
propagation, while the fractures of the latter condi-
specimens fabricated via L-PBF and LP-DED AM
tion were governed mainly by the nucleation,
processes. In addition, the effect of different heat-
growth, and coalescence of voids as well as final
Microstructure and Deformation Behavior of Additively Manufactured 17–4 Stainless Steel:
Laser Powder Bed Fusion vs. Laser Powder Directed Energy Deposition

Fig. 9. Tensile fracture surfaces of the L-PBF and LP-DED 17-4 PH SS specimens: (a) CA-H900, and (b) CA-H1150.

treatment procedures on the crystallographic tex- ferrite, with Widmanstätten ferrite grains dec-
ture and room-temperature tensile deformation orating the grain boundaries. The variation in
behavior of L-PBF and LP-DED specimens were the cooling/solidification rates induced by the L-
investigated. The following conclusions are drawn: PBF and LP-DED processes and the differences
in Creq/Nieq values for the L-PBF and LP-DED
1. The NHT L-PBF 17–4 PH SS specimen had a specimens may have caused these differences in
more refined microstructure, constituted of their NHT microstructures.
equiaxed ferrite grains and lath martensite, 2. The d-ferrite phase was retained in the
whereas the microstructure of the LP-DED microstructure of the LP-DED 17–4 PH SS
counterpart was composed of coarse massive
Nezhadfar, Gradl, Shao, and Shamsaei

specimen at room temperature. This was as- REFERENCES


cribed to the lower cooling/solidification rate in 1. N. Shamsaei, A. Yadollahi, L. Bian, and S.M. Thompson,
the LP-DED process and higher Creq/Nieq value Addit. Manuf. 8, 12 (2015).
of the LP-DED specimen compared to the L-PBF 2. D.D. Gu, W. Meiners, K. Wissenbach, and R. Poprawe, Int.
one, which stabilized ferrite in the microstruc- Mater. Rev. 57, 133 (2013).
3. T.D. Ngo, A. Kashani, G. Imbalzano, K.T.Q. Nguyen, and D.
ture. Hui, Compos. Part B 143, 172 (2018).
3. The over-aging heat treatment procedure (i.e., 4. T.F. Babuska, B.A. Krick, D.F. Susan, and A.B. Kustas,
CA-H1150) likely resulted in the reversion of Manuf. Lett. 28, 30 (2021).
austenite from martensite. However, the re- 5. A. Jinoop, C. Paul, and K. Bindra, J. Mater. Des. Appl. 233,
verted austenite was retained in the microstruc- 2376 (2019).
6. R. Molaei, A. Fatemi, N. Sanaei, J. Pegues, N. Shamsaei, S.
ture of the L-PBF specimen, while it was Shao, P. Li, D.H. Warner, and N. Phan, Int. J. Fatigue 132, 1
transformed to martensite in the LP-DED coun- (2020).
terpart upon cooling. The finer microstructure 7. J. Simpson, J. Haley, C. Cramer, O. Shafer, A. Elliott, B.
in the L-PBF specimen resulted in finer reverted Peter, L. Love, and R. Dehoff, ORNL/TM-2019-1190. Oak
Ridge Natl. Lab. 1190, 1 (2019).
austenite, which hindered the martensitic 8. Q. Guo, C. Zhao, M. Qu, L. Xiong, L.I. Escano, S.M.H. Hoj-
transformation. jatzadeh, N.D. Parab, K. Fezzaa, W. Everhart, T. Sun, and
4. A full heat treatment cycle is required for the L- L. Chen, Addit. Manuf. 28, 600 (2019).
PBF and LP-DED 17–4 PH SS specimens to 9. A. Saboori, A. Aversa, G. Marchese, S. Biamino, M. Lom-
diminish their variation in texture. Both the L- bardi, and P. Fino, Appl. Sci. 10, 3310 (2020).
10. M. Ma, Z. Wang, and X. Zeng, Mater. Sci. Eng. A 685, 265
PBF and LP-DED specimens had a twin-based (2017).
martensite with a K-S relationship to the parent 11. K.N. Amato, S.M. Gaytan, L.E. Murr, E. Martinez, P.W.
austenite. Shindo, J. Hernandez, S. Collins, and F. Medina, Acta Ma-
5. The L-PBF 17–4 PH SS specimen outperformed ter. 60, 2229 (2012).
12. F. Liu, X. Lin, H. Leng, J. Cao, Q. Liu, C. Huang, and W.
the LP-DED and wrought counterparts with the Huang, Opt. Laser Technol. 45, 330 (2013).
same heat treatment in the room-temperature 13. L.L. Parimi, G. Ravi, D. Clark, and M.M. Attallah, Mater.
tensile properties. This was attributed to the Charact. 89, 102 (2014).
finer microstructure of the L-PBF specimens. 14. W. Xu, M. Brandt, S. Sun, J. Elambasseril, Q. Liu, K. La-
6. The d-ferrite phase in the CA-H900 heat treated tham, K. Xia, and M. Qian, Acta Mater. 85, 74 (2015).
15. B.E. Carroll, T.A. Palmer, and A.M. Beese, Acta Mater. 87,
LP-DED specimen reduces the ductility com- 309 (2015).
pared with that of the L-PBF counterpart even 16. J. Schneider, JOM 72, 1085 (2020).
though the LP-DED specimen had lower tensile 17. C. Doñate-Buendia, R. Streubel, P. Kürnsteiner, M.B.
strength. The high contrast between the Schmid Wilms, F. Stern, J. Tenkamp, E. Bruder, S. Barcikowski, B.
Gault, K. Durst, J.H. Schleifenbaum, F. Walther, and B.
factor of the d-ferrite phase with the martensite Gökce, Procedia CIRP 94, 41 (2020).
matrix in the {110}<111> slip system (i.e., high 18. M.R. Stoudt, R.E. Ricker, E.A. Lass, and L.E. Levine, JOM
incompatibility with matrix), as well as the 69, 506 (2017).
sharp edges of the d-ferrite phase, justified the 19. S. Cheruvathur, E.A. Lass, and C.E. Campbell, JOM 68, 930
reduction in the ductility of the material. (2015).
20. W. Liu, J. Ma, M.M. Atabaki, R. Pillai, B. Kumar, U.
7. The quasi-cleavage facets on the fracture sur- Vasudevan, H. Sreshta, and R. Kovacevic, Lasers Manuf.
faces of the CA-H900 treated specimens repre- Mater. Process 2, 74 (2015).
sented brittle behavior of the L-PBF and LP- 21. S. Cao, Determination of the Fe-Cr-Ni and Fe-Cr-Mo phase
DED 17–4 PH SS in the CA-H900 condition, diagrams at intermediate temperatures using a novel dual-
anneal diffusion-multiple approach, Ph.D thesis, The Ohio
while the fibrous fracture surface with dimples State University, 2013.
depicted ductile behavior of the material in CA- 22. TCFE9 Thermo-Calc Software, 1 (2017).
H1150 condition. 23. P.D. Nezhadfar, E. Burford, K. Anderson-Wedge, B. Zhang,
S. Shao, S.R. Daniewicz, and N. Shamsaei, Int. J. Fatigue
123, 168 (2019).
ACKNOWLEDGEMENTS 24. P.D. Nezhadfar, P. Gradl, S. Shuai, and N. Shamsaei, In
This paper is based upon the work partially fun- Proc. 32nd Annu. Int. Solid Free. Fabr. Symp. – An Addit.
Manuf. Conf. (2021).
ded by the National Aeronautics and Space 25. ASTM Standard E8/E8M-13a, Standard Test Methods for
Administration (NASA) under Award Tension Testing of Metallic Materials (2006).
#80MSFC19C0010. Any subjective views or opin- 26. T. LeBrun, T. Nakamoto, K. Horikawa, and H. Kobayashi,
ions that might be expressed in the paper do not Mater. Des. 81, 44 (2015).
necessarily represent the views of NASA or the 27. L. Facchini, N. Vicente, I. Lonardelli, E. Magalini, P. Ro-
botti, and A. Molinari, Adv. Eng. Mater. 12, 184 (2010).
United States Government. 28. M. Alnajjar, F. Christien, K. Wolski, and C. Bosch, Addit.
Manuf. 25, 187 (2019).
29. P.D. Nezhadfar, K. Anderson-Wedge, S.R. Daniewicz, N.
CONFLICT OF INTEREST Phan, S. Shao, and N. Shamsaei, Addit. Manuf. 36, 101604
(2020).
On behalf of all authors, the corresponding author 30. R.K. Okagawa, R.D. Dixon, and D.L. Olson, Weld. Res.
states that there is no conflict of interest. Suppl. 62, 204s (1983).
31. P.D. Nezhadfar, R. Shrestha, N. Phan, and N. Shamsaei,
Int. J. Fatigue 124, 188 (2019).
Microstructure and Deformation Behavior of Additively Manufactured 17–4 Stainless Steel:
Laser Powder Bed Fusion vs. Laser Powder Directed Energy Deposition
32. C. Celada-Casero, J. Sietsma, and M.J. Santofimia, Mater. 37. ASTM A693-16, Standard specification for precipitation-
Des. 167, 107625 (2019). hardening stainless and heat-resisting steel plate, sheet, and
33. D.H. Ping, S.Q. Guo, M. Imura, X. Liu, T. Ohmura, M. strip (2016).
Ohnuma, X. Lu, T. Abe, and H. Onodera, Sci. Rep. 8, 14264 38. M.E. Stevenson, P.D. Umberger, and S.F. Uchneat, Fracture
(2018). appearance and mechanisms of deformation and fracture
34. C. Cayron, Acta Crystallogr. Sect. A 69, 498 (2012). (ASM International, 2021).
35. P.D. Nezhadfar, A. Zarei-Hanzaki, S.S. Sohn, and H.R.
Abedi, Mater. Sci. Eng. A 665, 10 (2016). Publisher’s Note Springer Nature remains neutral with re-
36. G.E. Dieter and D.J. Bacon, Mechanical metallurgy
gard to jurisdictional claims in published maps and institutional
(McGraw-Hill, New York, 1986).
affiliations.

You might also like